Next Article in Journal
Actions of Novel Angiotensin Receptor Blocking Drugs, Bisartans, Relevant for COVID-19 Therapy: Biased Agonism at Angiotensin Receptors and the Beneficial Effects of Neprilysin in the Renin Angiotensin System
Next Article in Special Issue
Design of an Internal/External Bicontinuous Conductive Network for High-Performance Asymmetrical Supercapacitors
Previous Article in Journal
Correction: Peters et al. Benchmarking the ACEnano Toolbox for Characterisation of Nanoparticle Size and Concentration by Interlaboratory Comparison. Molecules 2021, 26, 5315
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Flower-like Highly Open-Structured Binder-Free Zn-Co-Oxide Nanosheet for High-Performance Supercapacitor Electrodes

1
Department of Intelligent Manufacturing, Yibin University, Yibin 644000, China
2
Metallurgy and Materials Engineering Department, Dawood University of Engineering and Technology, Karachi 74800, Pakistan
3
Department of Physics and Bei**g Key Laboratory of Energy Conversion and Storage Materials, Bei**g Normal University, Bei**g 100084, China
4
Department of Chemistry, Faculty of Science, King Khalid University, Abha 61413, Saudi Arabia
5
Departamento de Quimica Organica, Universidad de Cordoba, E14014 Cordoba, Spain
6
School of Physical Science and Technology, Lanzhou University, Lanzhou 730000, China
7
Unité Transformations & Agroressources, Univ Artois—UniLaSalle, F-62408 Béthune, France
8
Dipartimento di Ingegneria dell’Innovazione, Università del Salento, 73100 Lecce, Italy
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(15), 4850; https://doi.org/10.3390/molecules27154850
Submission received: 16 June 2022 / Revised: 26 July 2022 / Accepted: 27 July 2022 / Published: 29 July 2022
(This article belongs to the Special Issue Porous Materials for Electrochemical Energy Storage and Conversion)

Abstract

:
Scientific research is being compelled to develop highly efficient and cost-effective energy-storing devices such as supercapacitors (SCs). The practical use of SC devices is hindered by their low energy density and poor rate capability due to the binding agents in fabricating electrodes. Herein, we proposed flower-like highly open-structured binder-free ZnCo2O4 micro-flowers composed of nanosheets supported in nickel foam (ZnCoO@NF) with improved rate capability up to 91.8% when current varied from 2 to 20 A·g−1. The ZnCoO@NF electrode exhibited a superior specific capacitance of 1132 F·g−1 at 2 A·g−1 and revealed 99% cycling stability after 7000 cycles at a high current density of 20 A·g−1. The improved performance of the ZnCoO@NF electrode is attributed to the highly stable structure of the micro/nano-multiscale architecture, which provides both the high conduction of electrons and fast ionic transportation paths simultaneously.

1. Introduction

Clean and renewable energy sources are in high demand due to current energy shortages and environmental concerns [1,2]. Supercapacitors (SCs) with a high power density and long cycle life with minimum pollution have attracted considerable interest [3] because of their intriguing advantages, for example, lightweight and outstanding power, which are in high demand in manufacturing smartphones and handy gadgets [4]. SCs usually store electrical charge on the electrode material’s surface by adsorption/desorption, known as electrochemical double-layer capacitors (EDLCs). In the second type, the near-surface area via reduction/oxidation states changes, and a reversible redox reaction occurs; this charge-storing mechanism is known as a pseudocapacitive mechanism [5]. Pseudocapacitors (PCs) have a higher capacitance and energy density than EDLC-based SCs, but it is necessary to upgrade PCs before they can be used in real-world applications [6]. The redox reactions governed by the electrode material’s multiple oxidation states significantly impact the performance of PCs [7]. Designing nanostructured electrode materials for PC with high redox-active sites and improved performance is challenging. In this context, interchanging more than one electron for every redox reaction is considered to be advantageous. The performance of PCs can be improved by preparing the multiple valance state materials with abundant electroactive sites [8]. Transition metal oxides have sparked interest in PCs over the last decade due to easy fabrication and cost-effectiveness [4]. Binary metal oxides possess more valance states and higher electrical conductivity than single metal oxides, up to two times higher, and exhibit better performance [9]. The poor performance of transition metal oxides may be due to their inability to conduct electricity. Following prior research, increasing the electrical conductivity of host oxides by do** them with metals or semi-conductor elements is essential for achieving high power density in SCs [10]. Thus, Zn do** significantly increases the electrical conductivity of ZnCo2O4, especially in comparison with Co3O4, and redox-inert Zn2+ offers good synergistic impacts and considerably improves the rate-performance of cobalt-oxide.
Cobalt-based binary metal-oxide electrode materials such as ZnCo2O4, FeCo2O4, CuCo2O4, NiCo2O4, and MnCo2O4 can significantly increase the performance of PCs [11,12,13]. It is believed that ZnCo2O4 is the most promising candidate due to the low toxicity of zinc metal and reasonably significant capacitance of Co3O4 (3560 F·g−1), and it also possesses good stability at low cost [14]. However, the rate performance of cobalt-based binary metal-oxides is poor, and for SCs, the rate–performance parameter is critical. To enhance the ZnCo2O4 electrode’s overall performance, three general strategies are typically employed: (i) development of nanostructured materials with a visually appealing structure and a large surface area; (ii) research and development of new composites with high conductivity, which includes carbon nanotubes and graphene; and (iii) enhancement of the SC device’s potential window, as energy density is inversely proportional to the potential window’s square (E = 1/2CV2) [15]. Increasing the capacitance and voltage of SCs can improve their energy density. Capacitance can be enhanced by increasing the electrode material’s surface area and pore size [16]. Binders prevent active materials from falling off during electrode operation by cohering the electrode material with substrates. When making an electrode, the binder should be able to provide the necessary strength and pore sizes. However, binding or cohesive agents inevitably cover some of the active material’s surface area or pores. As a result, the electrochemical performance of SCs is directly influenced by the binders and their content in the electrodes [17]. Binder-free electrodes have recently been found to significantly increase electrical conductivity and ionic diffusion paths and thus improve overall performance. In addition, the “dead volume” in the electrode material can be avoided using a binder-free electrode with a highly porous and robust construction [18].
Saravanakumar et al. fabricated a binder-free hierarchical electrode of ZnCo2O4, showing a suitable specific capacitance of 92 mF·cm−2 at 0.5 mA·cm−2 [19]. Fu et al. made a binder-free flower-like ZnCo2O4 electrode with a good capacitance of 689.4 F·g−1 at 1 A·g−1 with excellent cyclic stability of 97.1% after 1500 cycles [20]. Kamble et al. used ZnCo2O4 to make binder-free electrodes and obtained a good capacitance of 127.8 F·g−1 at 1 mA·cm−2 with excellent cycling stability of 3000 cycles [21]. Although considerable efforts to increase the capabilities and cycling stability of ZnCo2O4 have been ongoing, the cycling stability and rate performance are still crucial and need to be significantly improved. The electrochemical performance of ZnCo2O4 electrodes can be further improved by exploring the various morphologies. We propose that electrodes of ZnCo2O4 with micro/nano morphology can increase surface area and pore size, resulting in increased energy density and rate performance.
Herein, we proposed binder-free ZnCo2O4 micro-flowers composed of nanosheets supported on (ZnCoO@NF), improving the rate capability up to 91.8% when current varied from 2 to 20 A·g−1. The ZnCoO@NF electrode exhibits a superior capacitance of 1132 F·g−1 at 2 A·g−1 with excellent cycling stability of ~99% after 7000 cycles at a high current density of 20 A·g−1. The improved performance of the ZnCoO@NF electrode is attributed to the highly stable structure of micro/nano-architecture, which provides both high electronic conduction and fast ionic transportation paths.

2. Materials and Methods

2.1. Zn-Co-Oxide Synthesis on Nickel Foam

The 2D ZnCo-oxide micro-flowers assembled by nanosheets on nickel foam (NF) were grown directly using a low-cost and straightforward hydrothermal synthesis method. In a usual cycle, a homogenous solution was made by mixing 2 mmol of Zn(NO3)2·6H2O, 4 mmol of Co(NO3)2·6H2O, 2 mmol of NH4F, and 6 mmol urea in 40 mL of deionized water and ultra-sonicating the solution for 60 min. Afterwards, a hygienic piece of nickel foam (NF) was hung on the inside walls of a Teflon autoclave, and the solution mixture was injected into the Teflon autoclave. After securing the autoclave, it was placed in an electric oven and heated to 135 °C for 16 h. The NF was assembled with ZnCoO-oxide precursors, washed with deionized water and ethanol, and sonicated for 5 min to remove the remaining particles from the surface of NF. For further processing, the ZnCoO@NF was dehydrated in an air furnace for 12 h. To improve the crystallinity, the ZnCoO@NF precursor was heated for 2 h at 250 °C in the air atmosphere. The experiment was repeated twice and yielded the same results each time.

2.2. Fabrication of Electrode

The as-synthesized ZnCoO@NF can be directly used as an electrode for SCs without any binder. The ZnCoO@NF is cut into 1 × 1 cm2 pieces and directly used as a working electrode. Based on the difference in the weights of the bare NF and ZnCoO@NF, 1.26 mg·cm−2 of active material was found on the surface of NF.

2.3. Characterization of Material

Field emission scanning electron microscopy (SEM, FEI Nova 400, FEI Company, Hillsboro, OR, USA) and transmission electron microscopy (TEM, JEM 2100F TEM, Jeol, Tokyo, Japan) were both used to examine the materials’ microstructures and morphology. X-ray powder diffraction (XRD) (PANalyticalX’Pert Powder, Malvern Panalytical, Malvern, UK) was used to determine the crystal structure of the synthesized product. The X-ray source was an Escalab (250**) (Thermo Fisher Scientific, Waltham, MA, USA) with an Al Ka (1486.5 eV). The specific surface area calculation was made with a Quantachrome Instrument (Version 5.12, Boynton Beach, FL, USA) and the Brunauer-Emmett-Teller (BET) method.

2.4. Electrochemical Tests Measurements

The electrochemical tests, including cyclic voltammetry (CV), galvanostatic charge/discharge (GCD), and electrochemical impedance spectroscopy (EIS), were performed in an open-circuit position in a frequency ranging from 0.001 Hz to 100 kHz. All these tests were performed using a platinum sheet (counter electrode), Ag/AgCl (reference electrode), and ZnCoO@NF (working electrode) in a 6 KOH electrolyte at room temperature using an electrochemical workstation (CHI 660E, CH Instruments, Wuhan, China) in a three-electrode configuration. The specific capacitance (Cs) was calculated using the following Equation (1):
Cs = I × Δ t d m × Δ V
where Cs (F·g−1) represents the specific capacitances, I represents the current (A), Δtd (s) represents the discharging time, m (g) represents the active mass of the ZnCoO material on NF, and the potential window is represented by ΔV (V).

3. Results

Figure 1 gives a schematic route for the fabrication process of ZnCoO on nickel foam. In order to improve the mechanical and crystalline, the ZnCoO was post-annealed for the short term at 250 °C for 2 h in an air atmosphere with a ramp rate of 2 °C·min−1.
Figure 2a–c shows the morphology of as-prepared ZnCoO@NF at different magnifications, reflecting the 2D micro-flowers made by nanosheets. Figure 2a shows a low-resolution SEM image, demonstrating the rough surface with the growth of high-density ZnCoO nanosheet architecture on the entire surface of the NF and reflecting a compact design and interconnected structure. The high-magnification SEM images show that the various nanosheets make a micro-flower-like structure (Figure 2b). An open channel structure formed by interconnected nanosheets and an average thickness of 38 nm is shown in Figure 2c. Compared and a pure nanosheet, these micro-flower patterns increase the surface area of the electrode and enhance the electrode/electrolyte interface, making it better suited for use in SCs.
TEM analysis was further conducted to investigate the ZnCoO structure and surface morphologies in more depth. A characteristic TEM image with low resolution represents nanosheets of ZnCoO (Figure 3a,b). Figure 3a shows that a 2D structure can be derived from the interconnected nanosheets, increasing the intrinsic electronic conductivity. Figure 3b shows that the lattice fringe spacing is 0.23 nm, corresponding to the 111 planes of ZnCo2O4 (PDF# 23-1390). These findings are also in line with the results of the SEM. It should be noted that in the SEM and TEM images, we can see some residual nanoparticles. It is confirmed that there is no effect from these residual nanoparticles on the electrochemical performance of the electrode.
Figure 4a shows the specific surface area and pore size distribution (PSD) of the ZnCoO@NF electrode recorded at 77.5 K. BET absorbed volume of ZnCoO@NF is about 50 cm3/g at 1.0 P/P0, which means that ZnCoO@NF has a highly porous structure (55 m2·g−1). Moreover, this study demonstrated an H1-type hysteresis loop, which suggested the presence of a mesoporous structure in the solution. The PSD was calculated from the desorption isotherm using the Barrett-Joyner-Halenda (BJH) method, as shown in Figure 4b. The PSD of ZnCoO@NF is centred at about 9.6 and 61 nm, reinforcing the prepared sample’s mesoporous structure with pore sizes ranging from 9.6 to 61 nm. These findings show that the ZnCoO@NF sample has a greater surface area that leads to efficient ion and electron ship**, implying that high electrochemical activity in the ZnCoO@NF electrode is expected. For efficient PC electrodes, a high specific surface area can effectively enhance the electrochemical performance and provide a large number of electroactive sites for PCs.
The XRD pattern of ZnCoO@NF (Figure 5a) is well-indexed according to the standard spinel phase ZnCo2O4 (JCPDF# 23-1390). The particular diffraction peaks of ZnC2O4 located at 21°, 32°, 37°, 39°, 44.7°, 56°, 59.5°, and 66° correspond, respectively, to 111, 220, 311, 222, 422, 511, and 440 lattice planes. Compared with the standard ZnCo2O4 spinel, these results are consistent with the previous findings [20]. The lattice plane of 400 at 44.7° is overlapped with the XRD peak of NF. No other impurities are detected except two intense peaks from the NF at 45° and 52.5° of 2θ, demonstrating the high purity of the sample. The JCPDF# 23-1390 of pure Co3O4 was added as a comparison to confirm the difference between ZnCo2O4 and Co3O4. The diffraction peaks of as-prepared ZnCoO@NF clearly matched with JCPDF# 23-1390 and confirm the pure phase of ZnCo2O4. The crystal structure ZnCo2O4 nanostructures are shown in Figure 5b. Binary spinel metal oxide ZnCo2O4 with Co2+ at the tetrahedral sites (8a) in the spinel Co3O4 replaced by Zn2+ offers a better stability. In addition, the spinal structure can help increase the redox reaction sites, which boosts the specific capacitance of bimetallic oxide-based electrodes. Thus, choosing the bimetallic oxide and crystal structure is critical for overall performance and structural stability during the electrochemical discharge process [22].
The Raman spectrum further characterized the as-synthesized ZnCoO@NF to confirm the Raman band characteristics. From Figure 5c, the Raman spectrum of ZnCoO@NF demonstrates four active vibrational bands located at 183, 472, 513, and 682 cm−1 assigned to the characteristic F2g, Eg, F2g, and F2g phonon modes, respectively. The presence of Zn, Co, and O in these distinct peaks of the Raman spectrograms suggests the formation of ZnCo2O4. This indicates that the prepared sample is free of contaminants [23].
The electrochemical performance results of ZnCoO@NF are shown in Figure 6. The CV was performed at various scan rates in a potential window ranging from 0.0 to 0.5 V. The CV curves indicate two redox peaks at approximately 0.38 and 0.42 V (vs Ag/AgCl), which represent the typical pseudocapacitive behavior of the ZnCoO@NF associated with Co3+/Co2+ during the charge/discharge process [19]. The reduction peak shift is related to the internal resistance of the electrode materials, which further confirms the pseudocapacitive features with enhanced charge storage [24]. Moreover, the CV possesses a large area under the curve with high current response, indicating better charge storage behavior. The electrodes’ excellent redox reversibility can be seen in the redox peaks caused by reactions involving Co–O/Co–O–OH [25].
The electrochemical property of the ZnCoO@NF electrode was tested based on charging and discharging at various current densities to accurately estimate the charge storage capability. The GCD measurements of the ZnCoO@NF electrode were carried out in a potential window ranging from 0.0 to 0.45 V. Figure 6 depicts typical GCD curves of ZnCoO@NF at various current densities (2 to 20 A·g−1) (b). The GCD curves are nearly symmetric (Coulombic efficiency 98%), have apparent redox behavior, and agree well with the CV results, demonstrating the pseudocapacitive charge storage features. The charge and discharge curves primarily have voltage terraces at about 0.3 V and 0.36 V, respectively. Equation (1) can be used to calculate the specific capacitance Cs based on the curves. Based on Equation (1), the obtained specific capacitances of the ZnCoO@NF electrode are about 1132, 1120, 1100, 1060, and 1040 F·g−1 at current densities of 2, 3, 5, 10, and 20 A·g−1, respectively, as shown in Figure 6c. When the discharge current density changes from 2 to 20 A·g−1, the capacitance remains at 91.8%. The capacitance of the ZnCoO@NF electrode is much better than other electrodes such as Gd-doped CeOx nanoflowers (280 F·g−1 in 1 M NaOH at 1 V·s−1) [26] and mesoporous Co3O4 nanosheets on carbon foam (106 F·g−1 in 1 M NaOH) at a scan rate of 0.1 V·s−1 [27].
Because of its multi-channel porous nature and enhanced ionic transportation, the ZnCoO@NF demonstrated exceptional specific capacitance, far exceeding that of previously reported ZnCo2O4-based electrode materials.
The total capacitance of ZnCoO@NF can be attributed to the following significant contributions: (1) the micro-flowers composed of nanosheets with highly open structures supply abundant active sites for the Faraday reaction and offers easy diffusion paths for fast ionic transportation; (2) the electrochemical performance can be promoted by improving the specific area, pore size, and multi-channel nano/micro-architecture.
Furthermore, electrochemical impedance spectroscopy (EIS) was conducted to study the kinetics of charge transfer in the ZnCoO@NF electrode to confirm the charge transport properties. Figure 6d displays the Nyquist plots of the EIS spectra for the ZnCoO@NF electrode. The diameter of the semi-circle is used to estimate the charge transfer resistance (Rct = 1.52 Ω), which means that ion diffusion and electrolyte penetration in the active materials are reduced. The solution resistance (Rs = 0.13 Ω) is the combination of the electrolyte’s and electrode’s internal resistance. The values of Rct and Rs are smaller than in pristine Co3O4 (Rs = 1.32 Ω) [28] and ZnO (Rs = 2.82 Ω and (Rct = 2.25 Ω) [29]. The high-sloped curve at the high-frequency region indicates the fast ionic diffusion with enhanced electroactive sites in the ZnCoO@NF electrode material. The Nyquist plot was fitted with an equivalent circuit using ZSim view software, as presented in the inset of Figure 6d. The equivalent circuit shows Rs, Rct, C (the double layer capacitance), and W (the Warburg diffusion element). Hence, the ZnCoO@NF electrode possesses more capacitive-type charge-storage features [30].
Another vital parameter to consider when evaluating the practical applications of SCs is their cycling stability. Thus, the cycling stability test of the ZnCoO@NF electrode was performed by deliberately running it for 7000 charge/discharge cycles at a high current density of 20 A·g−1, and the results are shown in Figure 7a. After 7000 cycles, the specific capacitance remains at 99%. In the long-term cycling test, the ZnCoO@NF exhibited excellent charge stability and high performance, as shown by the obtained results. Moreover, at 20 A·g−1, the first and last five GCD cycles are shown in Figure 7b, indicating that the ZnCoO@NF electrode exhibits excellent stability before and after testing. Due to the large exposed surface area of its unique micro-flower composed of nanosheet architectures, the ZnCoO@NF electrode has impressive electrochemical performance. Micro-flowers can be grown directly onto the substrate without a binder, enabling direct electron intercalation between the substrate and the active material. It is easier to conduct a fast ion-diffusion reaction with the active material due to the micropores in the NF substrate. Because of their high rate capability, excellent capacitance, and admirable cycling stability, ZnCoO@NF electrodes are potential candidates for pseudocapacitive SC applications. The detailed comparisons of the current work with the previous literature in terms of electrode material, capacitance, current density, number of cycles, and retention are presented in Table 1.

4. Conclusions

The ZnCoO@NF was successfully synthesized using a straightforward hydrothermal method followed by a thermal procedure. The ZnCoO@NF electrode exhibits excellent electrochemical properties in an aqueous KOH electrolyte. The ZnCoO@NF electrode demonstrates a high specific capacitance of 1132 F·g−1 at 2 A·g−1 and 1040 F·g−1 at 20 A·g−1 with improved rate capability up to 91.8% when the current increased from 2 to 20 A·g−1. The ZnCoO@NF electrode shows excellent cycling stability of ~99% after 7000 cycles at a high current density of 20 A·g−1. Future energy storage devices will benefit from the new binder-less electrode material with high specific capacitance and rate capability. In light of the promising results, SCs combined with other energy conversion/storage devices may be commercially available soon.

Author Contributions

Conceptualization, M.S.J. and Q.A.; methodology, Q.A.; software, A.M.; validation, M.S.J.; formal analysis, Q.A., N.J.; investigation, M.S.J.; resources, S.H.S.; data curation, M.S.J. and Q.A.; writing—original draft preparation, S.H.S.; writing—review and editing, A.M., P.B., P.M., A.A., M.A.B.; visualization, P.B., P.M., N.J.; supervision, M.S.J.; project administration, M.A.B. and P.B.; funding acquisition, P.M., A.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Deanship of Scientific Research at King Khalid University for funding through the general research program R.G.P. 2/196/43.

Data Availability Statement

The data presented in this study are available on request from the first author Qasim Abbas, who is responsible for the performed experiments.

Acknowledgments

The authors acknowledge the scientific and technical input and support from the School of Physical Science and Technology, Lanzhou University, China. This work was financially supported by the scientific research start-up grant for Youth Researchers at Lanzhou University. The theoretical calculation was supported by the Supercomputing Center of Lanzhou University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kalair, A.; Abas, N.; Saleem, M.S.; Kalair, A.R.; Khan, N. Role of energy storage systems in energy transition from fossil fuels to renewables. Energy Storage 2021, 3, e135. [Google Scholar] [CrossRef] [Green Version]
  2. Barik, R.; Ingole, P.P. Challenges and prospects of metal sulfide materials for supercapacitors. Curr. Opin. Electrochem. 2020, 21, 327–334. [Google Scholar] [CrossRef]
  3. Dai, S.; Liu, Z.; Zhao, B.; Zeng, J.; Hu, H.; Zhang, Q.; Chen, D.; Qu, C.; Dang, D.; Liu, M. A high-performance supercapacitor electrode based on N-doped porous graphene. J. Power Sources 2018, 387, 43–48. [Google Scholar] [CrossRef] [Green Version]
  4. Javed, M.S.; Shaheen, N.; Hussain, S.; Li, J.; Shah, S.S.A.; Abbas, Y.; Ahmad, M.A.; Raza, R.; Mai, W. An ultra-high energy density flexible asymmetric supercapacitor based on hierarchical fabric decorated with 2D bimetallic oxide nanosheets and MOF-derived porous carbon polyhedra. J. Mater. Chem. A 2019, 7, 946–957. [Google Scholar] [CrossRef]
  5. Li, S.; Yu, C.; Yang, J.; Zhao, C.; Zhang, M.; Huang, H.; Liu, Z.; Guo, W.; Qiu, J. A superhydrophilic “nanoglue” for stabilizing metal hydroxides onto carbon materials for high-energy and ultralong-life asymmetric supercapacitors. Energy Environ. Sci. 2017, 10, 1958–1965. [Google Scholar] [CrossRef]
  6. Sun, S.; Zhai, T.; Liang, C.; Savilov, S.V.; ** and UV irradiation. RSC Adv. 2014, 4, 35067–35071. [Google Scholar] [CrossRef]
  7. Xu, Z.; Younis, A.; Chu, D.; Ao, Z.; Xu, H.; Li, S. Electrodeposition of Mesoporous Co3O4 Nanosheets on Carbon Foam for High Performance Supercapacitors. J. Nanomater. 2014, 2014, 902730. [Google Scholar] [CrossRef] [Green Version]
  8. Siyal, S.H.; Javed, M.S.; Ahmad, A.; Sajjad, M.; Batool, S.; Khan, A.J.; Akram, S.; Alothman, A.A.; Alshgari, R.A.; Najam, T. Free-standing 3D Co3O4@NF micro-flowers composed of porous ultra-long nanowires as an advanced cathode material for supercapacitor. Curr. Appl. Phys. 2021, 31, 221–227. [Google Scholar] [CrossRef]
  9. Abbas, Q.; Javed, M.S.; Ahmad, A.; Siyal, S.H.; Asim, I.; Luque, R.; Albaqami, M.D.; Tighezza, A.M.J.C. ZnO nano-flowers assembled on carbon fiber textile for high-performance supercapacitor’s electrode. Coatings 2021, 11, 1337. [Google Scholar] [CrossRef]
  10. Chen, H.; Jiang, J.; Zhao, Y.; Zhang, L.; Guo, D.; **a, D. One-pot synthesis of porous nickel cobalt sulphides: Tuning the composition for superior pseudocapacitance. J. Mater. Chem. A 2015, 3, 428–437. [Google Scholar] [CrossRef]
  11. **ao, X.; Wang, G.; Zhang, M.; Wang, Z.; Zhao, R.; Wang, Y. Electrochemical performance of mesoporous ZnCo2O4 nanosheets as an electrode material for supercapacitor. Ionics 2018, 24, 2435–2443. [Google Scholar] [CrossRef]
  12. Yu, H.; Zhao, H.; Wu, Y.; Chen, B.; Sun, J. Electrospun ZnCo2O4/C composite nanofibers with superior electrochemical performance for supercapacitor. J. Phys. Chem. Solids 2020, 140, 109385. [Google Scholar] [CrossRef]
  13. Moon, I.K.; Yoon, S.; Oh, J. Three-dimensional hierarchically mesoporous ZnCo2O4 nanowires grown on graphene/sponge foam for high-performance, flexible, all-solid-state supercapacitors. Chem.–A Eur. J. 2017, 23, 597–604. [Google Scholar] [CrossRef] [PubMed]
  14. Chen, H.; Wang, J.; Han, X.; Liao, F.; Zhang, Y.; Gao, L.; Xu, C. Facile synthesis of mesoporous ZnCo2O4 hierarchical microspheres and their excellent supercapacitor performance. Ceram. Int. 2019, 45, 8577–8584. [Google Scholar] [CrossRef]
  15. Rajesh, J.A.; Min, B.-K.; Kim, J.-H.; Kang, S.-H.; Kim, H.; Ahn, K.-S. Facile hydrothermal synthesis and electrochemical supercapacitor performance of hierarchical coral-like ZnCo2O4 nanowires. J. Electroanal. Chem. 2017, 785, 48–57. [Google Scholar] [CrossRef]
  16. Liu, B.; Zhang, J.; Wang, X.; Chen, G.; Chen, D.; Zhou, C.; Shen, G. Hierarchical three-dimensional ZnCo2O4 nanowire arrays/carbon cloth anodes for a novel class of high-performance flexible lithium-ion batteries. Nano Lett. 2012, 12, 3005–3011. [Google Scholar] [CrossRef]
  17. Bai, W.; Tong, H.; Gao, Z.; Yue, S.; **ng, S.; Dong, S.; Shen, L.; He, J.; Zhang, X.; Liang, Y. Preparation of ZnCo2O4 nanoflowers on a 3D carbon nanotube/nitrogen-doped graphene film and its electrochemical capacitance. J. Mater. Chem. A 2015, 3, 21891–21898. [Google Scholar] [CrossRef]
  18. Zhou, G.; Zhu, J.; Chen, Y.; Mei, L.; Duan, X.; Zhang, G.; Chen, L.; Wang, T.; Lu, B.J.E.A. Simple method for the preparation of highly porous ZnCo2O4 nanotubes with enhanced electrochemical property for supercapacitor. Electrochim. Acta 2014, 123, 450–455. [Google Scholar] [CrossRef]
Figure 1. The schematic representation of the synthesis process of ZnCoO@NF through the hydrothermal method.
Figure 1. The schematic representation of the synthesis process of ZnCoO@NF through the hydrothermal method.
Molecules 27 04850 g001
Figure 2. The SEM images of ZnCoO@NF at different magnifications: (a) low, (b) medium, and (c) high.
Figure 2. The SEM images of ZnCoO@NF at different magnifications: (a) low, (b) medium, and (c) high.
Molecules 27 04850 g002
Figure 3. Typical (a) low-resolution and (b) high-resolution TEM images of the ZnCoO@NF nanosheets.
Figure 3. Typical (a) low-resolution and (b) high-resolution TEM images of the ZnCoO@NF nanosheets.
Molecules 27 04850 g003
Figure 4. (a) N2 adsorption and desorption isotherms; (b) pore-size distribution of ZnCoO@NF.
Figure 4. (a) N2 adsorption and desorption isotherms; (b) pore-size distribution of ZnCoO@NF.
Molecules 27 04850 g004
Figure 5. Physical characterizations of ZnCoO@NF: (a) xrd pattern, (b) crystal structure (c) Raman spectra.
Figure 5. Physical characterizations of ZnCoO@NF: (a) xrd pattern, (b) crystal structure (c) Raman spectra.
Molecules 27 04850 g005
Figure 6. (a) CV curve at varied scan rates, (b) charge-discharge curve at current densities, (c) specific capacitance, and (d) electrochemical impedance spectra of the ZnCoO@NF electrode.
Figure 6. (a) CV curve at varied scan rates, (b) charge-discharge curve at current densities, (c) specific capacitance, and (d) electrochemical impedance spectra of the ZnCoO@NF electrode.
Molecules 27 04850 g006
Figure 7. (a) Cycling stability tests of ZnCoO@NF electrode (b) Five GCD cycles before and after stability test.
Figure 7. (a) Cycling stability tests of ZnCoO@NF electrode (b) Five GCD cycles before and after stability test.
Molecules 27 04850 g007
Table 1. Comparisons of the ZnCoO@NF electrode with previous literature in terms of electrode material, capacitance, current density, number of cycles, and retention.
Table 1. Comparisons of the ZnCoO@NF electrode with previous literature in terms of electrode material, capacitance, current density, number of cycles, and retention.
Serial No.Electrode/MaterialsCapacitance
(F·g−1)
Current Density
(A·g−1)
No. of Cycles
(n)
Retention (%)References
1ZnCoO@NF11322700099This work
2ZnCo2O4865.351.0100073[31]
3ZnCo2O4/C composite90.18.0100096[32]
4ZnCo2O4 NWs/rGO1116.62500093.4[33]
5ZnCo2O4-microsphere6891.0500098[34]
6ZnCo2O4 nanowires6942.0200085[35]
7ZnCo2O4 nanoflowers77010300089[36]
8ZnCo2O4 microspheres6471.0300096[23]
9Flower-like ZnCo2O46891.0150097[37]
10ZnCo2O4 nanoparticles 7101.0300084[38]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Abbas, Q.; Siyal, S.H.; Mateen, A.; Bajaber, M.A.; Ahmad, A.; Javed, M.S.; Martin, P.; Joly, N.; Bocchetta, P. Flower-like Highly Open-Structured Binder-Free Zn-Co-Oxide Nanosheet for High-Performance Supercapacitor Electrodes. Molecules 2022, 27, 4850. https://doi.org/10.3390/molecules27154850

AMA Style

Abbas Q, Siyal SH, Mateen A, Bajaber MA, Ahmad A, Javed MS, Martin P, Joly N, Bocchetta P. Flower-like Highly Open-Structured Binder-Free Zn-Co-Oxide Nanosheet for High-Performance Supercapacitor Electrodes. Molecules. 2022; 27(15):4850. https://doi.org/10.3390/molecules27154850

Chicago/Turabian Style

Abbas, Qasim, Sajid Hussain Siyal, Abdul Mateen, Majed A. Bajaber, Awais Ahmad, Muhammad Sufyan Javed, Patrick Martin, Nicolas Joly, and Patrizia Bocchetta. 2022. "Flower-like Highly Open-Structured Binder-Free Zn-Co-Oxide Nanosheet for High-Performance Supercapacitor Electrodes" Molecules 27, no. 15: 4850. https://doi.org/10.3390/molecules27154850

Article Metrics

Back to TopTop